Next Article in Journal
Effects of Substrate Pulse Bias on Corrosion Behavior of Tetrahedral Amorphous Carbon Thin Films in Acidic and Chloride Solutions
Previous Article in Journal
Turbulence Effects on the Dynamic Characteristics of Non-Circular Journal Bearings Under Large Sommerfeld Number Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Experimentally Validated Cavitation Model for Hydrodynamic Bearings Using Non-Condensable Gas

1
Institute of Tribology and Energy Conversion Machinery, Clausthal University of Technology, 38678 Clausthal-Zellerfeld, Germany
2
Product Development, Ruhr-University Bochum, 44801 Bochum, Germany
3
Gleitlagertechnik Weißbacher GmbH, 46519 Alpen, Germany
*
Author to whom correspondence should be addressed.
Lubricants 2025, 13(4), 140; https://doi.org/10.3390/lubricants13040140
Submission received: 22 January 2025 / Revised: 27 February 2025 / Accepted: 20 March 2025 / Published: 25 March 2025

Abstract

:
Despite great research effort in recent decades, cavitation in hydrodynamic journal bearings is still a not completely understood phenomenon. In particular, it is unclear which proportions of different cavitation types are present in a bearing. Novel experimental results show a clear deviation from the predictions of hydrodynamic lubrication theory. This article presents a new approach for modeling cavitation in hydrodynamic bearings by using computational fluid dynamics with the volume of fluid method and a phase of non-condensable gas in the lubrication oil. The validation of the model is achieved through the simulation of a large Offset-Halves Bearing and a subsequent comparison of the results with various experimental data, including the fractional film content. In the results, cavitation also occurs in the convergent gap due to a pressure drop caused by inertia forces. The findings indicate that the cavitation effects in oil-lubricated hydrodynamic bearings are caused by a special form of gaseous cavitation, designated as pseudo-cavitation. The presented model with non-condensable gas is able to reproduce the observed phenomena excellently.

Graphical Abstract

1. Introduction

Cavitation is the formation of a gas or vapor phase as a result of a pressure drop. Cavitation is an effect that forms in many slide bearings, particularly in the area of the opening gap, which is why it occurs primarily in fixed-lobe bearings. Cavitation can impair operational safety. The state of knowledge about cavitation in hydrodynamic lubrication was reviewed by Braun and Hannon [1]. Three different types of cavitation can be differentiated:
  • Vaporous cavitation occurs when the pressure falls below the vapor pressure of the liquid phase and boiling begins. Depending on the temperature, Ref. [1] gives values for the vapor pressure p v of typical oil mixture components between 2.1 · 10 4 Pa and 52.4 kPa that are close to the vacuum.
  • Gaseous cavitation is caused by dissolved gas that will be desorbed when pressure drops below saturation pressure. Braun and Hannon provide values for the temperature-dependent saturation pressure p s of between 34.48 kPa and 101.37 kPa .
  • Pseudo-cavitation is a special case of gaseous cavitation when the gas is completely undissolved. This type of cavitation is caused by the expansion of dispersed gas bubbles due to depressurization. The processes of molecular dissolution and absorption do not occur, and the total amount of free gas is constant ( R des = R abs = 0 ). The gas can enter the system either via the lubricant or directly from the environment.
In Figure 1, the mass transfer between the single phases is illustrated for the general case of cavitation. In the event that the pressure falls below the saturation pressure, the dissolved gas desorbs from the oil at a desorption rate R des . Conversely, if the saturation pressure is exceeded, the gas is absorbed at an absorption rate R abs . Similarly, if the pressure falls below the vapor pressure, the oil vaporizes at a vaporization rate R vap . Conversely, if the vapor pressure is exceeded, the vapor condenses at an absorption rate R con .
The most common mathematical description of cavitation in hydrodynamic lubrication is based on the Reynolds equation with Elrod’s cavitation algorithm [2]. Elrod extends the Reynolds equation by an additional variable θ , denoted as fractional film content or density ratio, and ( 1 θ ) is named the void fraction.
· ρ θ h 3 12 η p = · ( ρ θ U h ) + t ( ρ θ h )
For closure, he introduces the complementary condition
p > p cav θ = 1 + 1 β ( p p cav ) , p = p cav 0 θ < 1 ,
where β is the compressibility factor resulting in a value of θ slightly greater than one. The influence of a gas phase on the density ratio is neglected in this consideration. This algorithm is mass conservative and gives a pressure distribution that aligns closely with experimental results but does not differentiate between the cavitation types and does not consider the real physical processes in the cavitation zone. However, it assumes a specific cavitation formation with gas regions and liquid streamers. The lubricant is advected through the streamers by the Couette flow, and the Poiseuille flow is assumed to be zero. Various modifications and different types of implementations have been made to this algorithm.
Models for vaporous cavitation have been derived by simplifying the Rayleigh–Plesset equation, for example, by Schnerr and Sauer [3], Singhal [4], and Zwart et al. [5]. These models are typically used in combination with a numerical solution of the Navier–Stokes equation and are available in several computational fluid dynamic software packages. The Singhal model is also able to take non-condensable gas into account.
Peeken and Benner [6] were the first to describe a gaseous cavitation model based on a saturation equilibrium. Another equilibrium model was developed by a research group at Tsinghua University in Beijing: Li et al. [7] present a gaseous cavitation model based on a saturation equilibrium, which they apply to a journal bearing in combination with the Reynolds equation. Hao et al. [8] extended the model of Li et al. for a non-saturation equilibrium and applied it in combination with the Reynolds equation to a thrust bearing. Song et al. [9] used a similar gaseous cavitation model to Li et al., which they applied to various slide bearings in combination with the Reynolds equation. The outcomes of these studies demonstrate analogous results to those observed in the context of ordinary cavitation boundary conditions. Ding et al. [10,11] also present a similar gaseous cavitation model and apply it to a journal tilting pad bearing in a computational fluid dynamics simulation, finding a better agreement with experimental data when using the non-equilibrium model than the equilibrium model.
Ibata et al. [12] present an empirical gaseous cavitation model considering the effects of dynamic stimulation on the saturation pressure. Osterland et al. [13] present a gaseous cavitation model that they adapt to hydraulic oil together with the vapor cavitation model by Zwart et al. using experimental data. They set the absorption rate to zero. Reinke et al. [14] investigated gaseous cavitation with high-speed digital photography in a Couette–Poiseuille flow for kerosene oil, finding that the dissolved gas accumulates in the oil over time, resulting in increasing pseudo-cavitation until all the gas is undissolved. They adapted their results to an empirical model and set the absorption rate to zero. In a previous publication [15], the authors of this paper presented a high-speed journal bearing with reduced pad size. A novel methodology for quantifying the fractional film content in a large journal bearing with local resolution was proposed and applied for the first time in [16], where it was tested on this bearing. The results have shown clear deviations to the predictions made by the classical lubrication theory. Comparing Figure 17 in [15] with Figures 13–18 in [16], the experimental results reveal the presence of large partially flooded areas, especially behind the oil feed and at the sides of the pads. In contrast, the lubrication theory, using the Reynolds equation and Elrod’s cavitation algorithm, predict an almost fully-flooded gap. Only at the end of the loaded pad, where the gap narrows, is cavitation predicted in agreement with the measurement. The visualization of cavitation zones had already been presented qualitatively by previous authors using optical methods, for example, in [17], which exhibits cavitation formations similar to those observed in [16], such as the formation of a large fillament in the divergent region. The present study demonstrates that this unpredicted cavitation effect is the result of a combination of inertia forces and undissolved gases. For this purpose, a pseudo-cavitation model and its validation by comparing numerical results with comprehensive experimental data are presented below.

2. Mathematical Model

2.1. Governing Equations

For describing the flow of the given problem, a computational fluid dynamics model that is based on Eulerian treatment by the three-dimensional Navier–Stokes equation is employed. The volume of fluid (VOF) approach is applied to describe the multiphase nature of the flow. With this approach, each phase α within a specified volume V takes up a partial volume V α . Based on this assumption, the volume fraction r α = V α / V is defined as a continuous field. Overall, the volume conservation equation
α r α = 1
must be fulfilled in the entire solution domain. To maintain mass conservation, a continuity equation is balanced separately for each phase.
r α ρ α t + · r α ρ α v = 0 .
Moreover, the conservation of momentum and energy is balanced for all phases as a whole, which is known as a homogeneous model. The momentum conservation equation reads
ρ v t + · ρ v v = · τ p .
The thermal energy equation is
ρ h t + · ( ρ v h ) = · λ T + Φ ,
whereby h is the static enthalpy with d h = c p d T and Φ = τ : v is the dissipation term. In the homogenous model, the mixture fluid properties are assumed to be volume-weighted averages, accordingly ρ = α r α ρ α for the density, η = α r α η α for the dynamic viscosity, and λ = α r α λ α for the specific heat conductivity. Additional turbulence and transition effects are considered by using the Shear Stress Transport (SST) turbulence model [18] together with the Gamma-Theta transition model [19]. For the solid domain, the energy equation simplifies in a way that the dissipation term is zero and the solid velocity is specified. Solid deformation is considered using a Lagrangian treatment by a coupled structural mechanic simulation that is based on the principle of virtual displacements in
Ω σ : δ ε dV = Ω t · δ u dA + Ω ρ k · δ u dV
with the external stress t and volumetric forces ρ k .

2.2. Material Laws

When modeling gaseous cavitation with non-condensable gases, the compressibility of the gas phase emerges as the pivotal property. Therefore, the ideal gas law is used as the equation of state to determine the density as
ρ = R T p .
The viscous stress tensor related to a Newtonian fluid is
τ = η v + v T 2 3 · v 1 ,
where 1 is the unity tensor. The formula by Falz [20] is used to describe a temperature-dependent viscosity with the reference viscosity η 0 , the reference temperature T 0 , and the Falz exponent l.
η = η 0 T T 0 l
The density and specific heat of the fluid are described as linear-dependent on the temperature, see (13) and (14). For the Cauchy stress tensor, a linear isotropic model is used.
σ = E 1 + ν ε D + E 3 ( 1 2 ν ) Tr ( ε ) + 3 a Δ T 1 ,
wherein Tr ( ε ) means the trace and ε D the deviator of the deformation tensor.

2.3. Numerical Solution

The governing equations are approximated using methods of weighted residuals, the finite element method is employed for the structural model, and the element-based finite volume method is used for the fluid model and the solid heat transfer. A hybrid structured/unstructured technique employing 8-node hexahedron and 6-node prism elements is applied for the mesh generation. The solution of the fluid model and the conjugated heat transfer, as implemented in the Ansys CFX solver, is coupled with an external structural mechanics solver in a custom framework. The fluid model is solved for a few hundreds of iterations, with a time step size between 10 5 and 10 4 s. Subsequently, the temperature and pressure fields are utilized as a load for the structural mechanics model. The mounted bearing is elastically supported in a steel housing, which is modeled as a hollow cylinder with a large outer diameter, the mesh of which is not shown here. The rotor position is determined in order to reach a specified operation point using an algebraic model, whereby tolerances of ± 0.05 MPa for the specific load and ± 0 . 5 for the load angle is targeted in this study. Consequently, the structural model generates a mesh displacement, which is then utilized to update the meshes for the fluid model. The submodel’s rotor displacement, fluid model, and solid model are solved repeatedly to obtain a stationary converged solution. The solution can be initialized with that of a previous operating point. The computation time depends on the operating point. A typical time is about two days using 4 partitions.

3. Test Case

3.1. Geometry

As a test case (see Figure 2), an Offset-Halves Bearing is used, which was previously investigated in [15,16]. More details about the bearing and the experimental setup can be found in these references. The bearing has a nominal diameter of 500 mm , an outer diameter of 800 mm , and a width of 350 mm . To reduce the power loss, the width of the upper sliding surface is smaller. Both sliding surfaces are separated by two deep pockets with 142 mm length. The oil supply is applied by two leading edge grooves (LEG) with a chord length of 30 mm with common ring channel feeding. A hydrostatic jacking pocket that is closed by a check valve during static operation with a chord length of 84 mm , a width of 80 mm , and a radius of 90 mm is located adjacent to the load direction. The geometric parameters are listed in Table 1.
The related computational grids are displayed in Figure 3. The lubrication gap is meshed with 20 nodes across its height. The model utilizes a total of 2 · 10 5 fluid nodes and 6 · 10 4 solid nodes. Mesh Independency is verified in Appendix A.

3.2. Material Parameters

The shaft and the bearing liner are machined from carbon steel. A 4 mm thick coating of white metal is applied to act as a sliding surface. The corresponding material parameters are listed in Table 2.
Table 3 contains the material properties of the used ISO VG 32 turbine oil. applied for lubrication in this study.
The density and the specific heat capacity of the lubricant can be described in close approximation as linearly dependent on temperature. A fit to tabular values given in Table 3 yields the following empirical correlations:
ρ = 837.1 kg m 3 0.64 kg m 3 K ( T 50   ° C )
c p = 2140.2 J kg K + 3.5832 J kg K 2 ( T 50   ° C ) .
An adjustment of Equation (10) to these data results in η 0 = 35.28 mPa s and l = 1.79 . The reference temperature used is T 0 = 40 °C. The density of air at a reference state T = 25 °C, p = 101 , 325 Pa is assumed by ρ = 1.185 kg / m 3 ; the specific heat at constant pressure of air is c p = 1004.4 J / ( kg K ) ; and the viscosity and the heat conductivity are set to be η = 1.83 · 10 5 Pa s and λ = 0.0261 W / ( m K ) .

3.3. Boundary Conditions

At the inlet on the outside of the ring channel, an oil volume flow rate of 450 L / min is fed with an inlet temperature of 50   ° C . The volume fraction of the gas phase at the inlet is determined by
r g = α B α B + p p 0 T 0 T .
with the Bunsen solubility coefficient α B = 0.09 . At the outer fluid boundary, atmospheric pressure and pure air ( r g = 1 ) are specified. The specific load is varied in the range of p ¯ = 1 MPa to p ¯ = 5 MPa and directed to the center of the jacking pocket defined at φ = 180 . The rotational speed of the rotor is varied from n = 1000 rpm to n = 3600 rpm in the positive z-direction. For the investigated operation points, mirror symmetry is assumed at z = 0 . The remaining boundaries are treated as adiabatic. These are the free surface of the fluid in the ring channel, the outer surface and face of the bearing, and the inner surface and face of the rotor.

4. Numerical Results

4.1. Pressure

Figure 4 depicts a comparison between measured and predicted hydrodynamic film pressure. The simulated pressure shows some irregularities at the end of the hydrostatic jacking pocket. These might be a numerical issue due to the sharp gradients at this location, although it could be physical and not captured by the sensor. Both the experiment and the simulation demonstrate that the pressure drops below the atmospheric pressure shortly behind the lubrication pocket. This is presented in detail in Figure 5 for experimental and simulation data. Here, it should be noted that the pressure sensor provides a smoothed signal due to its size and delay. This pressure drop can be explained by the acceleration of the fluid and the resulting inertia forces. This is illustrated in the section in Figure 5b.

4.2. Volume Fraction and Fractional Film Content

Figure 6a shows the liquid volume fraction r l of the simulation in the entire fluid region for the maximum investigated rotor speed and 4 MPa specific load. In Figure 6b, the liquid volume fraction at the bearing center is depicted across the circumferential angle φ and the distance y from the rotor surface across the gap. It can be seen that the volume fraction in the cavitation regions behind the oil supply varies across the lubrication gap in such a way that the oil phase separates partially at the sliding surface, while the volume fraction in the diverging gap zone is almost constant. Such a combination of streamers and adhered layers is typical for stationary loaded bearings.
The fractional film content for the simulation is evaluated as the average volume fraction of the lubricant across the gap height.
θ = 1 h 0 h r l d y
In Figure 7, the fractional film content is compared with the experimental results for different operation points. It can be seen that the cavities behind the leading edge grooves ➀ and ➁ occur at speeds above 2000 rpm . At lower speeds, only ambient air entering the gap from sides ➂ and ➃ is visible in the results.
This also indicates that the lubricating film pressure in this area is lower than the atmospheric pressure. With increasing speed, the pressure behind the oil feed pockets becomes lower and the expansion of the dispersed gas in the oil becomes more pronounced. With increasing load, these partially flooded and unflooded regions become smaller at the loaded pad and larger at the unloaded pad because of the changed film thickness. At high loads, the ordinary cavitation caused by a diverging gap at the end of the loaded pad can be observed ➄. The simulated and measured oil distributions are in overall agreement.
An alternative application of the gaseous cavitation model from Osterland et al. [13] resulted in an insufficient desorption rate. This supports the assumption that desorbed gas accumulates in the oil during operation and, consequently, pseudo cavitation occurs at stationary operation. Describing this shift from gaseous to pseudo cavitation would require the transient modeling of the full oil circuit, which would necessitate a considerable computational effort. Furthermore, simpler modeling approaches using a single phase of an oil–gas mixture with fixed composition (specified mass fractions) based on the Navier–Stokes as well as the Reynolds equation led to similar results to Elrod’s cavitation algorithm. In the investigated bearing, the mass fractions are not constant at all, especially in the areas that ingest enviromental air through the lateral boundary of the lubrication gap. It is essential to consider the multiphase nature of this cavitation effect by solving a single transport equation for each phase to ensure accurate modeling. Inertia forces must also be taken into account in order to describe the pressure drop.

4.3. Film Thickness

Figure 8 illustrates the film thickness in the bearing center along the circumferencial angle and along the axial direction at the position of the minimum film thickness. It can be seen in Figure 8b that the film thickness varies along the axial direction. This results from the thermal deformation of the rotor and bearing.

4.4. Temperature

The temperature within the bearing was measured by 123 type J (Fe-CuNi) thermocouples (see Figure 2). Figure 9 compares the measured temperature of the sensors in the bearing center with the predictions. These sensors are applied approximately 9 mm below the sliding surface. The accuracy is good at low to medium speed. At higher speeds, however, the predicted temperature is too low. This currently represents a limitation.

4.5. Power Loss

Figure 10 shows a satisfactory agreement between calculated and calorimetrically measured frictional power loss. However, at high speed some variance exists, which potentially originates from turbulence modeling. Prediction is particularly difficult at speeds of 3000 rpm and 3600 rpm at low load, as a large area of the lubricating film is in the laminar-turbulent transition.

4.6. Effect of the Bunsen Solubility Coefficient

The Bunsen solubility coefficient specifies the volume of dissolved gas per volume of liquid. For mineral oil, a range of α B = 0.08 0.1 is given in the literature. A larger amount of air may be present if free air is induced into the oil during operation and if the oil is contaminated or aged. In order to validate the chosen value of α B for the present model, it was varied and compared in terms of agreement of the fractional film content with the measurements. The results are shown in Figure 11. It can be seen that the higher the α B chosen, the narrower the area of incoming air on the sides, especially of the unloaded sliding surfaces due to the higher compressibility. Therefore, the correspondence is better for the investigated operating point, which indicates the presence of additional free gas.

5. Conclusions

In this paper, a thermal-elastohydrodynamic model of a large hydrodynamic Offset-Halves Bearing using non-condensable gases for describing cavitation is presented and validated with comprehensive experimental data.
By the utilization of the presented computational fluid dynamics numerical model, it is shown that cavitation for the given test bearing can be sufficiently described by non-condensable gases and that the cavitation type is pseudo cavitation. This is because undissolved gas accumulates in the lubricating oil over time due to the low resorption rate. As a result, common gaseous cavitation only occurs at the beginning of operation and only pseudo-cavitation is observed during stationary operation. The cavities occur wherever the pressure drops below atmospheric pressure, but not only in the divergent gap. In contrast, the described cavitation effect cannot be explained by the use of the classical hydrodynamic lubrication theory. The pressure drop that occurs subsequent to the oil feed is attributable to an inertia effect, which arises from the acceleration of the lubricant. A similar effect is known at the leading edge of tilting pad bearings where the lubricant is decelerated and a ram pressure occurs [21]. The pressure drop results in an expansion of the gas phase that separates from the liquid phase. These cavities as well as the pressure drop, which are not predicted by the thin film theory, are reproduced very accurately by our model. This shows that cavitation in hydrodynamic bearings is more complex than previously assumed and is not fully understood. In particular, the influence of inertia forces on the dynamic pressure must be taken into account. The partially flooded areas resulting from gaseous cavitation can potentially compromise the operational safety especially, at high speed and low load operation. Further investigation is necessary to elucidate the effects on rotor dynamics.

Author Contributions

Conceptualization, methodology, validation, and investigation, S.W., A.E., T.H., M.S. and C.W.; software, S.W.; writing—original draft, S.W., A.E. and T.H.; writing—review and editing, M.S., C.W., H.S. and B.B.; visualization, S.W.; and supervision, funding acquisition, C.W., B.B. and H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the German Federal Ministry of Economic Affairs and Climate Action. The financial support was assigned by the Industrial Research Association (AiF e. V.) in project number IGF 21389 N.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

Author Christoph Weißbacher was employed by the company Gleitlagertechnik Weißbacher GmbH. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Nomenclature

Latin SymbolsGreek Symbols
aThermal expansion rate α B Bunsen solubility coefficient
c p Specific heat at constant pressure β Compressibility factor
EYoung’s modulus ε Deformation tensor
hFilm thickness, static enthalpy η Dynamic viscosity
k External field θ Fractional film content
lFalz exponent λ Heat conductivity
m ˙ α β Mass transfer rate from phase α to β ν Poisson’s ratio
nRotational speed ρ Density
pAbsolute pressure σ Cauchy stress tensor
p cav Cavitation pressure τ Viscous stress tensor
p s Saturation pressure Φ Dissipation term
p v Vapor pressure φ Circumferencial angle
p ¯ Specific load Ω Integration domain
RGas constant Ω Boundary of Ω
R abs Absorption rate
R abs Condensation rateOperators
R des Desorption rate ( ) : ( ) Frobenius scalar product
R vap Vaporisation rate ( ) · ( ) Dot product
rVolume fraction ( ) Gradient
tTime · ( ) Divergence
t External stress ( ) ( ) Dyadic product
TTemperature ( ) D Deviator
U Surface velocity Tr ( ) Trace
u Solid displacement ( ) T Transposition
VVolume
v Fluid velocitySubscripts
x , y , z Spacial coordinates ( ) g Gas
( ) l Liquid
( ) v Vapor
( ) 0 Reference value

Appendix A. Mesh Independency Study

Proofing complete grid independence in such a simulation is intricate. However, this section demonstrates that a mesh refinement of the lubrication gap has no significant effect on the hydrodynamic pressure. Furthermore, empirical values regarding the mesh size from reference [15] can be used, whereby the present mesh is significantly finer.
For this study, the rotor position and the deformation are fixed. The mesh of the lubrication gap has two parameters regarding its fineness: the number of nodes across the gap (a value of 20 was used for the results) and a global element size that controlls the mesh size in the remaining directions (a value of 5 mm ). As illustrated in Figure A1a, with more than 15 nodes across the gap, the pressure just sligthly differs. Figure A1b demonstrates that even a substantial coarsening of the global element size has little effect on the overall pressure shape.
Figure A1. Effect of mesh refinement of the lubrication gap on the hydrodynamic pressure (a) variation of the number of nodes across the gap and (b) variation of the global element size.
Figure A1. Effect of mesh refinement of the lubrication gap on the hydrodynamic pressure (a) variation of the number of nodes across the gap and (b) variation of the global element size.
Lubricants 13 00140 g0a1

References

  1. Braun, M.J.; Hannon, W.M. Cavitation formation and modeling for fluid film bearings: A review. Proc. Inst. Mech. Eng. Part J J. Eng. Tribol. 2010, 224, 839–863. [Google Scholar] [CrossRef]
  2. Elrod, H.G. A cavitation algorithm. J. Lubr. Technol. 1981, 103, 350–354. [Google Scholar] [CrossRef]
  3. Schnerr, G.; Sauer, J. Physical and Numerical Modeling of Unsteady Cavitation Dynamics. In Proceedings of the 4th International Conference on Multiphase Flow, New Orleans, LA, USA, 27 May–1 June 2001. [Google Scholar]
  4. Singhal, A. Mathematical Basis and Validation of the Full Cavitation Model. J. Fluids Eng. 2002, 124, 617–624. [Google Scholar] [CrossRef]
  5. Zwart, P.; Gerber, A.G.; Belamri, T. A two-phase flow model for predicting cavitation dynamics. In Proceedings of the ICMF 2004 International Conference on Multiphase Flow, Yokohama, Japan, 30 May–3 June 2004. [Google Scholar]
  6. Peeken, H.; Benner, J. Beeinträchtigung des Druckaufbaus in Gleitlagern durch Schmierstoffverschäumung. VDI-Berichte 1985, 549, 373–397. [Google Scholar]
  7. Li, X.; Song, Y.; Hao, Z.; Gu, C. Cavitation Mechanism of Oil-Film Bearing and Development of a New Gaseous Cavitation Model Based on Air Solubility. J. Tribol. 2012, 134, 31701. [Google Scholar] [CrossRef]
  8. Hao, Z.; Gu, C. Numerical modeling for gaseous cavitation of oil film and non-equilibrium dissolution effects in thrust bearings. Tribol. Int. 2014, 78, 14–26. [Google Scholar] [CrossRef]
  9. Song, Y.; Gu, C.; Ren, X. Development and validation of a gaseous cavitation model for hydrodynamic lubrication. Proc. Inst. Mech. Eng. Part J J. Eng. Tribol. 2015, 2299, 1227–1238. [Google Scholar] [CrossRef]
  10. Ding, A.; Li, X.; Li, Y. Improvement and analysis for a gaseous cavitation model applied in a tilting pad journal bearing. In Proceedings of the ASME Turbo Expo 2019: Turbomachinery Technical Conference and Exposition, Phoenix, AZ, USA, 17–21 June 2019; Volume 7B: Structures and Dynamics. [Google Scholar] [CrossRef]
  11. Ding, A.; Ren, X.; Li, X.; Gu, C. A new gaseous cavitation model in a tilting-pad journal bearing. Sci. Prog. 2021, 104, 368504211029431. [Google Scholar] [CrossRef] [PubMed]
  12. Ibata, Y.; Matsuura, Y.; Takahashi, S.; Iga, Y. Development of gaseous cavitation model in hydraulic oil flow considering the effect of dynamic stimulation. IOP Conf. Ser. Earth Environ. Sci. 2019, 240, 62041. [Google Scholar] [CrossRef]
  13. Osterland, S.; Günther, L.; Weber, J. Experiments and CFD on vapor and gas cavitation for oil hydraulics. Chem. Eng. Technol. 2022, 46, 147–157. [Google Scholar] [CrossRef]
  14. Reinke, P.; Ahlrichs, J.; Beckmann, T.; Schmidt, M. High-Speed Digital Photography of Gaseous Cavitation in a Narrow Gap Flow. Fluids 2022, 7, 159. [Google Scholar] [CrossRef]
  15. Hagemann, T.; Vetter, D.; Wettmarshausen, S.; Stottrop, M.; Engels, A.; Weißbacher, C.; Bender, B.; Schwarze, H. A Design for High-Speed Journal Bearings with Reduced Pad Size and Improved Efficiency. Lubricants 2022, 10, 313. [Google Scholar] [CrossRef]
  16. Engels, A.; Wettmarshausen, S.; Stottrop, M.; Hagemann, T.; Weißbacher, C.; Schwarze, H.; Bender, B. Experimental Identification of the Void Fraction in a Large Hydrodynamic Offset Halves Bearing. Lubricants 2024, 13, 7. [Google Scholar] [CrossRef]
  17. Dowson, D.; Godet, M.; Taylor, C.M. Cavitation and Related Phenomena in Lubrication: Proceedings of the 1st Leeds-Lyon Symposium on Tribology, Held in The Institute of Tribology, Department of Mechanical Engineering, The University of Leeds, England, September 1974; Mechanical Engineering Publications Limited: London, UK, 1975. [Google Scholar]
  18. Menter, F.R. Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications. AIAA J. 1994, 32, 1598–1605. [Google Scholar] [CrossRef]
  19. Langtry, R.B. A Correlation-Based Transition Model Using Local Variables for Unstructured Parallelized CFD Codes. Doctoral Dissertation, Universität Stuttgart, Fakultät Energie, Verfahrens und Biotechnik, Stuttgart, Germany, 2006. [Google Scholar] [CrossRef]
  20. Falz, E. Grundzüge der Schmiertechnik, 2nd ed.; Springer: Berlin/Heidelberg, Germany, 1931. [Google Scholar]
  21. Hagemann, T.; Zeh, C.; Prölß, M.; Schwarze, H. The Impact of Convective Fluid Inertia Forces on Operation of Tilting-Pad Journal Bearings. Int. J. Rotating Mach. 2017, 2017, 5683763. [Google Scholar] [CrossRef]
Figure 1. Interphase mass transfer for the different types of cavitation.
Figure 1. Interphase mass transfer for the different types of cavitation.
Lubricants 13 00140 g001
Figure 2. General view of the test bearing and sectional views of the pockets.
Figure 2. General view of the test bearing and sectional views of the pockets.
Lubricants 13 00140 g002
Figure 3. Computational grids of (a) fluid, (b) bearing, and (c) rotor.
Figure 3. Computational grids of (a) fluid, (b) bearing, and (c) rotor.
Lubricants 13 00140 g003
Figure 4. Comparison of pressure between measurement and simulation (a) for the entire sliding surface, (b) the bearing center, and (c) the circumferential position of the maximum pressure.
Figure 4. Comparison of pressure between measurement and simulation (a) for the entire sliding surface, (b) the bearing center, and (c) the circumferential position of the maximum pressure.
Lubricants 13 00140 g004
Figure 5. Pressure drop behind the leading edge groove of the loaded pad at the bearing center (a) at n = 2000 rpm , p ¯ = 1.00 MPa and (b) at n = 3600 rpm , p ¯ = 1.00 MPa .
Figure 5. Pressure drop behind the leading edge groove of the loaded pad at the bearing center (a) at n = 2000 rpm , p ¯ = 1.00 MPa and (b) at n = 3600 rpm , p ¯ = 1.00 MPa .
Lubricants 13 00140 g005
Figure 6. Oil volume fraction of the simulation (a) at the entire bearing and (b) at the bearing center ( n = 3600 rpm , p ¯ = 4 MPa ).
Figure 6. Oil volume fraction of the simulation (a) at the entire bearing and (b) at the bearing center ( n = 3600 rpm , p ¯ = 4 MPa ).
Lubricants 13 00140 g006
Figure 7. Comparison of the fractional film content θ between measurement and simulation with α B = 0.09 for various operating points.
Figure 7. Comparison of the fractional film content θ between measurement and simulation with α B = 0.09 for various operating points.
Lubricants 13 00140 g007
Figure 8. Comparison of the film thickness (a) at the bearing center and (b) at the circumferential position of the minimum film thickness.
Figure 8. Comparison of the film thickness (a) at the bearing center and (b) at the circumferential position of the minimum film thickness.
Lubricants 13 00140 g008
Figure 9. Comparison of sensor temperature at the bearing center between measurement and simulation for three different operating points.
Figure 9. Comparison of sensor temperature at the bearing center between measurement and simulation for three different operating points.
Lubricants 13 00140 g009
Figure 10. Comparison of the frictional power loss between measurement and simulation for different operating points.
Figure 10. Comparison of the frictional power loss between measurement and simulation for different operating points.
Lubricants 13 00140 g010
Figure 11. Comparison of the fractional film content θ between measurement and simulation ( n = 3600 rpm , p ¯ = 4 MPa ).
Figure 11. Comparison of the fractional film content θ between measurement and simulation ( n = 3600 rpm , p ¯ = 4 MPa ).
Lubricants 13 00140 g011aLubricants 13 00140 g011b
Table 1. Geometric parameters of the test bearing.
Table 1. Geometric parameters of the test bearing.
ParameterNominalMeasuredUnit
Rel. bearing clearance 1 . 20 + 0.09 1.19
Preload unloaded pad 1 . 75 0.12 + 0.07 1.60
Preload loaded pad 1 . 75 0.12 + 0.07 1.73
Angle of curvature center of unloaded pad 235 247
Angle of curvature center of loaded pad 55 58
Bore diameter 500 . 470 + 0.044 mm
Pad radius 250 . 460 + 0.022 mm
Outer diameter800 mm
Width unloaded pad350 mm
Width loaded pad210 mm
Shaft outer diameter 499.870 mm
Shaft inner diameter150 mm
Table 2. Material parameters of the bearing.
Table 2. Material parameters of the bearing.
ParameterSteelWhite MetalUnit
Density ρ 78707400 kg / m 3
Young’s modulus E 205 · 10 9 56.5 · 10 9 N / m 2
Poisson’s ratio ν 0.280.331
Thermal expansion rate a 11 · 10 6 23.4 · 10 6 1
Specific heat c p 440230 J / ( kg K )
Heat conductivity λ 51.9 35 W / ( m K )
Table 3. Material parameter of the lubricant.
Table 3. Material parameter of the lubricant.
ParameterValueUnit
Temperature T20406080100120140 ° C
Density ρ 856.1 843.5 830.8 818.0 805.1 792.1 kg / m 3
Dynamic viscosity η 89.14 35.15 17.52 10.19 6.62 4.65 mPa s
Specific heat c p 2032.7 2247.7 2319.3 2391.0 2462.7 J / ( kg K )
Heat conductivity λ 0.134 0.134 0.134 0.134 0.134 0.134 0.134 W / ( m K )
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wettmarshausen, S.; Engels, A.; Hagemann, T.; Stottrop, M.; Weißbacher, C.; Schwarze, H.; Bender, B. An Experimentally Validated Cavitation Model for Hydrodynamic Bearings Using Non-Condensable Gas. Lubricants 2025, 13, 140. https://doi.org/10.3390/lubricants13040140

AMA Style

Wettmarshausen S, Engels A, Hagemann T, Stottrop M, Weißbacher C, Schwarze H, Bender B. An Experimentally Validated Cavitation Model for Hydrodynamic Bearings Using Non-Condensable Gas. Lubricants. 2025; 13(4):140. https://doi.org/10.3390/lubricants13040140

Chicago/Turabian Style

Wettmarshausen, Sören, Alexander Engels, Thomas Hagemann, Michael Stottrop, Christoph Weißbacher, Hubert Schwarze, and Beate Bender. 2025. "An Experimentally Validated Cavitation Model for Hydrodynamic Bearings Using Non-Condensable Gas" Lubricants 13, no. 4: 140. https://doi.org/10.3390/lubricants13040140

APA Style

Wettmarshausen, S., Engels, A., Hagemann, T., Stottrop, M., Weißbacher, C., Schwarze, H., & Bender, B. (2025). An Experimentally Validated Cavitation Model for Hydrodynamic Bearings Using Non-Condensable Gas. Lubricants, 13(4), 140. https://doi.org/10.3390/lubricants13040140

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop